首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
本文合成了3个新钌(Ⅱ)配合物,[Ru(bpy)2(SB)](PF62、[Ru(bpy)(SB)2](PF62和[Ru(SB)3](PF62(bpy=2,2’-bipyridine,SB=4,5-diaza-9,9’-spirobifluorene),通过核磁和元素分析对配合物的结构进行了确定。[Ru(bpy)2(SB)](PF62通过X射线单晶衍射确认了结构。研究了配合物的光物理性能。结果表明[Ru(bpy)2(SB)](PF62在乙腈中的发桔红光,波长为606nm,量子产率约为0.0012。在同样条件下[Ru(bpy)(SB)2](PF62和[Ru(SB)3](PF62的发光非常微弱甚至几乎没有发光。还研究了这些配合物的电致化学发光性能。随着配体中SB含量的增加,发光的峰电压从1.36V增加到1.58V,相对发光强度从731降低到52。  相似文献   

2.
合成了3个钌髤配合物,[Ru(bpy)2(SB)](PF6)2、[Ru(bpy)(SB)2](PF6)2和[Ru(SB)3](PF6)2(bpy=2,2′-bipyridine,SB=4,5-diaza-9,9′-spirobifluorene),通过核磁和元素分析对配合物的结构进行了确定。[Ru(bpy)2(SB)](PF6)2通过X射线单晶衍射确认了结构。研究了配合物的光物理性能。结果表明[Ru(bpy)2(SB)](PF6)2在乙腈中的发桔红光,波长为606 nm,量子产率约为0.001 2。在同样条件下[Ru(bpy)(SB)2](PF6)2和[Ru(SB)3](PF6)2的发光非常微弱甚至几乎没有发光。还研究了这些配合物的电致化学发光性能。随着配体中SB含量的增加,发光的峰电压从1.36 V增加到1.58 V,相对发光强度从731降低到52。  相似文献   

3.
Two new bichromophoric ruthenium(II) complexes, [Ru(bpy)2(bpy‐CM)](PF6)2 and [Ru(bpy)2(bpy‐CM343)](PF6)2 (bpy=2,2′‐bipyridine, CM=coumarin) with appended coumarin ligands have been designed and synthesized. The energy‐transfer‐based sensing of esterase by the complexes has been studied by using UV/Vis and luminescence spectroscopic methods. The cytotoxicity and the cellular uptake of one of the complexes have also been investigated.  相似文献   

4.
Fast-atom bombardment (FAB) mass spectrometry in the negative ionization mode enables the sputtering into the gas phase of the ruthenium complexes [Ru(2,2′-bipyridine[bpy])2(2,5-bis) (pyrydil)pyrazine[dpp])](PF6)2; [Ru(bpy)2,(2,3dpp)](PF6)2;[Ru(bpy)2,(2,3-dpp-Me)]( PF6)3; and [Ru(bpy)2(?-2,3-dpp)]2 RuCl2(PF6)4 as intact radical anions. These data, combined with those avaiiable from the positive FAB spectra allow a full characterization of the analytes.  相似文献   

5.
The substitution behavior of the monodentate Cl ligand of a series of ruthenium(II) terpyridine complexes (terpyridine (tpy)=2,2′:6′,2′′-terpyridine) has been investigated. 1H NMR kinetic experiments of the dissociation of the chloro ligand in D2O for the complexes [Ru(tpy)(bpy)Cl]Cl ( 1 , bpy=2,2’-bipyridine) and [Ru(tpy)(dppz)Cl]Cl ( 2 , dppz=dipyrido[3,2-a:2′,3′-c]phenazine) as well as the binuclear complex [Ru(bpy)2(tpphz)Ru(tpy)Cl]Cl3 ( 3 b , tpphz=tetrapyrido[3,2-a:2′,3′-c:3′′,2′′-h:2′′′,3′′′-j]phenazine) were conducted, showing increased stability of the chloride ligand for compounds 2 and 3 due to the extended π-system. Compounds 1 – 5 ( 4 =[Ru(tbbpy)2(tpphz)Ru(tpy)Cl](PF6)3, 5 =[Ru(bpy)2(tpphz)Ru(tpy)(C3H8OS)/(H2O)](PF6)3, tbbpy=4,4′-di-tert-butyl-2,2′-bipyridine) are tested for their ability to run water oxidation catalysis (WOC) using cerium(IV) as sacrificial oxidant. The WOC experiments suggest that the stability of monodentate (chloride) ligand strongly correlates to catalytic performance, which follows the trend 1 > 2 > 5 ≥ 3 > 4 . This is also substantiated by quantum chemical calculations, which indicate a stronger binding for the chloride ligand based on the extended π-systems in compounds 2 and 3 . Additionally, a theoretical model of the mechanism of the oxygen evolution of compounds 1 and 2 is presented; this suggests no differences in the elementary steps of the catalytic cycle within the bpy to the dppz complex, thus suggesting that differences in the catalytic performance are indeed based on ligand stability. Due to the presence of a photosensitizer and a catalytic unit, binuclear complexes 3 and 4 were tested for photocatalytic water oxidation. The bridging ligand architecture, however, inhibits the effective electron-transfer cascade that would allow photocatalysis to run efficiently. The findings of this study can elucidate critical factors in catalyst design.  相似文献   

6.
The synthesis of tri-heteroleptic complex of Ru(II) with diimine ligands is describe. Ten compounds [Ru(R2bpy) (biq) (L)][PF6]2 (R = H, CH3); L = 2,2′-bipyridine (bpy), 4,4′-dimethyl-2,2′-bipyridine (Me2bpy), 2,2′-bipyrimidine (bpm), 2,2′-biisoquinoline (biiq), 1,10-phenanthroline (phen), dipyrido[3,2-c:2′,3′-e]pyridazine (taphen), 2,2′-biquinoline (biq), 6,7-dihydrodipyrido[2,3-b:3,2-j][1,10]-phenanthroline (dinapy), 2-(2[pyridyl)quinoline (pq), 1-(2-pyrimidyl)pyrazole] (pzpm), 2,2′-biimidazole (H2biim) are characterized by elemental analysis, electronic and 1H-NMR spectroscopy. The relative photosustitution rates of biq in MeCN are given at three temperatures.  相似文献   

7.
Stable nanoscale cross‐linked polymer micelles containing Ru complexes (Ru‐CMs) were prepared from monomethoxy[poly(ethylene glycol)]‐block‐poly(L ‐lysine) (MPEG‐PLys) and [(bpy)2Ru(fmbpy)](PF6)2 (bpy=bipyridine, fmbpy=5‐formy‐5′‐methyl‐2,2′‐bipyridine). To stabilize the micelles, bifunctional glutaraldehyde was used as a cross‐linker to react with the free amino groups of the PLys block. After that, the Ru‐CMs showed very good stability in common solvents. The Ru‐CMs showed photocatalytic activity and selectivity in the oxidation of sulfides that were as high as those of the well‐known [Ru(bpy)3(PF6)2] complex, because the micelles were swollen in the methanol–sulfide mixture. Moreover, because of the nanoscale size of the particles and their high stability, the Ru‐CM photocatalysts can be readily recovered by ultrafiltration and reused without loss of photocatalytic activity. This work highlights the potential of using cross‐linked micelles as a platform for developing highly efficient heterogeneous photocatalysts for a number of important organic transformations.  相似文献   

8.
The complexes [Ru(bpy)2(pyESO)](PF6)2 and [Os(bpy)2(pyESO)](PF6)2, in which bpy is 2,2′‐bipyridine and pyESO is 2‐((isopropylsulfinyl)ethyl)pyridine, were prepared and studied by 1H NMR, UV–visible and ultrafast transient absorption spectroscopy, as well as by electrochemical methods. Crystals suitable for X‐ray structural analysis were grown for [Ru(bpy)2(pyESO)](PF6)2. Cyclic voltammograms of both complexes provide evidence for S→O and O→S isomerization as these voltammograms are described by an ECEC (electrochemical‐chemical electrochemical‐chemical) mechanism in which isomerization follows Ru2+ oxidation and Ru3+ reduction. The S‐ and O‐bonded Ru3+/2+ couples appear at 1.30 and 0.76 V versus Ag/AgCl in propylene carbonate. For [Os(bpy)2(pyESO)](PF6)2, these couples appear at 0.97 and 0.32 V versus Ag/AgCl in acetonitrile, respectively. Charge‐transfer excitation of [Ru(bpy)2(pyESO)](PF6)2 results in a significant change in the absorption spectrum. The S‐bonded isomer of [Ru(bpy)2(pyESO)]2+ features a lowest energy absorption maximum at 390 nm and the O‐bonded isomer absorbs at 480 nm. The quantum yield of isomerization in [Ru(bpy)2(pyESO)]2+ was found to be 0.58 in propylene carbonate and 0.86 in dichloroethane solution. Femtosecond transient absorption spectroscopic measurements were collected for both complexes, revealing time constants of isomerizations of 81 ps (propylene carbonate) and 47 ps (dichloroethane) in [Ru(bpy)2(pyESO)]2+. These data and a model for the isomerizing complex are presented. A striking conclusion from this analysis is that expansion of the chelate ring by a single methylene leads to an increase in the isomerization time constant by nearly two orders of magnitude.  相似文献   

9.
A chemo‐sensor [Ru(bpy)2(bpy‐DPF)](PF6)2 ( 1 ) (bpy=2,2′‐bipyridine, bpy‐DPF=2,2′‐bipyridyl‐4,4′‐bis(N,N‐di(2‐picolyl))formylamide) for Cu2+ using di(2‐picolyl)amine (DPA) as the recognition group and a ruthenium(II) complex as the reporting group was synthesized and characterized successfully. It demonstrates a high selectivity and efficient signaling behavior only for Cu2+ with obvious red‐shifted MLCT (metal‐to‐ligand charge transfer transitions) absorptions and dramatic fluorescence quenching compared with Zn2+ and other metal ions.  相似文献   

10.
Syntheses and Structures of Bis(4,4′‐t‐butyl‐2,2′‐bipyridine) Ruthenium(II) Complexes with functional Derivatives of Tetramethyl‐bibenzimidazole [(tbbpy)2RuCl2] reacts with dinitro‐tetramethylbibenzimidazole ( A ) in DMF to form the complex [(tbbpy)2Ru( A )](PF6)2 ( 1a ) (tbbpy: bis(4,4′‐t‐butyl)‐2,2′bipyridine). Exchange of the two PF6? anions by a mixture of tetrafluor‐terephthalat/tetrafluor‐terephthalic acid results in the formation of 1b in which an extended hydrogen‐bonded network is formed. According to the 1H NMR spectra and X‐ray analyses of both 1a and 1b , the two nitro groups of the bibenzimidazole ligand are situated at the periphery of the complex in cis position to each other. Reduction of the nitro groups in 1a with SnCl2/HCl results in the corresponding diamino complex 2 which is a useful starting product for further functionalization reactions. Substitution of the two amino groups in 2 by bromide or iodide via Sandmeyer reaction results in the crystalline complexes [(tbbpy)2Ru( C )](PF6)2 and [(tbbpy)2Ru( D )](PF6)2 ( C : dibromo‐tetrabibenzimidazole, D : diiodo‐tetrabibenzimidazole). Furthermore, 2 readily reacts with 4‐t‐butyl‐salicylaldehyde or pyridine‐2‐carbaldehyde under formation of the corresponding Schiff base RuII complexes 5 and 6 . 1H NMR spectra show that the substituents (NH2, Br, I, azomethines) in 2 ‐ 6 are also situated in peripheral positions, cis to each other. The solid state structure of both 2 , and 3 , determined by X‐ray analyses confirm this structure. In addition, the X‐ray diffraction analyses of single crystals of the complexes [(tri‐t‐butyl‐terpy)(Cl)Ru( A )] ( 7 ) and [( A )PtCl2] ( 8 ) display also that the nitro groups in these complexes are in a cis‐arrangement.  相似文献   

11.
Based on the a ligand BDPPZ [(9a,13a‐dihydro‐4,5,9,14‐tetraaza‐benzo[b]triphenylene‐11‐yl)‐phenyl‐methanone] (1) and its polypyridyl hetero‐ and homoleptic Ru(II) metal complexes, [Ru(bpy)2L](PF6)2 (2), [Ru(phen)2L](PF6)2 (3), [Ru(dafo)2L](PF6)2 (4), [Ru(dcbpy)2L](PF6)2 (5) and [RuL3](PF6)2 (6) (where, L = ligand, bpy = 2,2′‐bipyridine, phen = 1,10‐phenantroline, dafo = 4,5‐diazafluoren‐9‐one and dcbpy = 3,3′‐dicarboxy‐2,2′‐bipyridine), have been synthesized and characterized by elemental analysis, UV–vis, FT‐IR, 1H and 13C‐NMR spectra (for ligand), molar conductivity measurements and X‐ray powder techniques. The electrochemical parameters of the substituted ligand and its polypyridyl hetero‐ and homoleptic Ru(II) metal complexes are reported by cyclic voltammetry. UV–vis spectroscopy is used to compare the differences between the conjugated π systems in this ligand and its Ru(II) metal complexes. The polypyridyl hetero‐ and homoleptic Ru(II) metal complexes also tested as catalysts for the formation of cyclic organic carbonates from carbon dioxide and liquid epoxides which served as both reactant and solvent. The results showed that the [Ru(L)3](PF6)2 (6) complex is more efficient than the other Ru(II) complexes for the formation of cyclic organic carbonates from carbon dioxide. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

12.
Several isomers are possible when N4-tetradentate ligands coordinate to form metal complexes. Here we report the synthesis and structural analysis of cis-β-{[1,6-di(2′-pyridyl)(2,5-dibenzyl-2,5-diazahexane)(1,2-benzoquinone diimine)]ruthenium(II)} formed exclusively from the β-precusor, β-{[1,6-di(2′-pyridyl)(2,5-dibenzyl-2,5-diazahexane) (dimethylsulfoxide)chloride] ruthenium(II)} hexaflourophosphate. Ruthenium(II) complexes synthesised from 1,6-di(2′-pyridyl)-2,5-dibenzyl-2,5-diazahexane, produce only two isomers which can be separated by recrystallisation into α- and β-[Ru(picenbz2)(dmso)Cl]PF6 (where picenbz2 is 1,6-di(2′-pyridyl)-2,5-dibenzyl-2,5-diazahexane). The distinctively different proton NMR spectra of the isomers are an especially convenient feature with which to assess separation. Isomeric structure of the precursor, α or β, is conserved upon coordination of a bidentate ligand, such as benzene-1,2-diamine, 4,5-dimethyl-benzene-1,2-diamine, naphthalene-2,3-diamine, 2,2′-bipyridine, 1,10-phenanthroline or dipyrido[3,2-d:2′3′-f] quinoxaline, to produce complexes of the type α- or β-[Ru(picenbz2)(bidentate)](PF6)2. The synthesis, separation and characterisation of the α- and β-precursors and the α- and β-[Ru(picenbz2)(bidentate)](PF6)2 complexes are reported. Moreover, the crystal structures have been determined for β-[Ru(picenbz2)(dmso)Cl]PF6.0.5H2O (C30H37N4O1.5F6PSClRu); it is triclinic, space group P 1, a?=?9.987, b?=?12.883, c?=?14.287?Å, α?=?72.11, β?=?78.65, γ?=?88.39° and Z?=?2 and β-[Ru(picenbz2)(bqdi)](PF6)2, (C34H38N6F12P2Ru) which is triclinic, space group P 1, with a?=?10.129, b?=?10.338, c?=?19.587?Å, α?=?104.42, β?=?93.36, γ?=?92.00° and Z?=?2. The structures were determined at room temperature and refined by least-squares methods to R?=?0.044 for 5109 and R?=?0.075 for 3057 non-zero diffractometer data, respectively, for the dmso and bqdi species above.  相似文献   

13.
A homogeneous visible light photoredox TEMPO‐mediated selective oxidation of primary alcohols to the corresponding carbonyl compounds was developed using molecular oxygen from air as the terminal oxidant. Ru(bpy)3(PF6)2 (bpy: bipyridyl) and Ir(dtb‐bpy)(ppy)2(PF6) (dtb‐bpy: 4,4′‐di‐tert‐butyl‐2,2′‐bipyridyl; ppy: 2‐phenylpyridine) were used as the sensitizers.  相似文献   

14.
Three Ru(II) complexes [Ru(bpy)2(1-IQTNH)](ClO4)2 (1), [Ru(bpy)2(2-QTNH)](ClO4)2 (2) and [Ru(bpy)2(3-IQTNH)](ClO4)2 (3) (bpy = 2,2′-bipyridine, 1-IQTNH = 6-(isoquinolin-1-yl)-1,3,5-triazine- 2,4-diamine, 2-QTNH = 6-(quinolin-2-yl)-1,3,5-triazine- 2,4-diamine, 3-IQTNH = 6-(isoquinolin-3-yl)-1,3,5-triazine-2,4-diamine) have been synthesized and characterized by elemental analysis, 1H NMR spectroscopy, electrospray ionization mass spectrometry and X-ray crystallography. The electrochemical and spectroscopic properties of the complexes differ from those of [Ru(bpy)3]2+ owing to the structural differences between the ligands and their complexes.  相似文献   

15.
In an effort to create a molecule that can absorb low energy visible or near‐infrared light for photochemotherapy (PCT), the new complexes [Ru(biq)2(dpb)](PF6)2 (1, biq = 2,2′‐biquinoline, dpb = 2,3‐bis(2‐pyridyl)benzoquinoxaline) and [(biq)2Ru(dpb)Re(CO)3Cl](PF6)2 (2) were synthesized and characterized. Complexes 1 and 2 were compared to [Ru(bpy)2(dpb)](PF6)2 (3, bpy = 2,2′‐bipyridine) and [Ru(biq)2(phen)](PF6)2 (4, phen = 1,10‐phenanthroline). Distortions around the metal and biq ligands were used to explain the exchange of one biq ligand in 4 upon irradiation. Complex 1, however, undergoes photoinduced dissociation of the dpb ligand rather than biq under analogous experimental conditions. Complex 3 is not photoactive, providing evidence that the biq ligands are crucial for ligand photodissociation in 1. The crystal structures of 1 and 4 are compared to explain the difference in photochemistry between the complexes. Complex 2 absorbs lower energy light than 1, but is photochemically inert although its crystal structure displays significant distortions. These results indicate that both the excited state electronic structure and steric bulk play key roles in bidentate photoinduced ligand dissociation. The present work also shows that it is possible to stabilize sterically hindered Ru(II) complexes by the addition of another metal, a property that may be useful for other applications.  相似文献   

16.
Subtle ligand modifications on RuII-polypyridyl complexes may result in different excited-state characteristics, which provides the opportunity to tune their photo-physicochemical properties and subsequently change their biological functions. Here, a DNA-targeting RuII-polypyridyl complex (named Ru1 ) with highly photosensitizing 3IL (intraligand) excited state was designed based on a classical DNA-intercalator [Ru(bpy)2(dppz)] ⋅ 2 PF6 by incorporation of the dppz (dipyrido[3,2-a:2′,3′-c]phenazine) ligand tethered with a pyrenyl group, which has four orders of magnitude higher potency than the model complex [Ru(bpy)2(dppz)] ⋅ 2 PF6 upon light irradiation. This study provides a facile strategy for the design of organelle-targeting RuII-polypyridyl complexes with dramatically improved photobiological activity.  相似文献   

17.
The condensation of 3-amino-1H-1,2,4-triazole with benzaldehyde and terephthalaldehyde provides the bidentate and tetradentate Schiff bases 1,2,4-triazolo-3-imino-benzene L1H and 1,4-bis(1,2,4-triazolo-3-imino)benzene L2H2, respectively. The well characterized Schiff bases were allowed to react with cis-Ru(bpy)2Cl2 · 2H2O. Isomers of the mononuclear complexes Ru(bpy)2L1]PF6 · NH4PF6 (1a, N4) and [Ru(bpy)2L1]PF6 · 0.5NH4PF6 (1b, N2), and the dinuclear Ru(II) complexes [Ru(bpy)2L2Ru(bpy)2](PF6)2 · NH4PF6 (2a, N4N4), [Ru(bpy)2L2Ru(bpy)2](PF6)2 · NH4PF6 · 2H2O (2b, N2N2) and [Ru(bpy)2L2Ru(bpy)2](PF6)3 · NH4PF6 (2c, Ru(II)-Ru(III)) were separated by column chromatography and characterized by their elemental analysis, FAB mass and spectral (IR, NMR, UV–Vis) data. The data obtained suggest that the ligands are bound to the metal centre via the N4 and N2 atoms of the triazole moiety along with the N (imine) atom. The complexes display metal-to-ligand charge-transfer (MLCT) transitions in the visible region from the dπ(RuII) → πL transition. Highly intense ligand-based π→π transitions are observed in the UV region. A dual emission occurs from the N2 and N2N2 isomers.  相似文献   

18.
Crystal structures of organometallic aqua complexes [Cp*RhIII(bpy)(OH2)]2+ ( 1 , Cp* = η5‐C5Me5, bpy = 2,2′‐bipyridine) and [Cp*RhIII(6,6′‐Me2bpy)(OH2)]2+ ( 2 , 6,6′‐Me2bpy = 6,6′‐dimethyl‐2,2′‐bipyridine) used as key catalysts in regioselective reduction of NAD+ analogues were determined definitely by X‐ray analysis. The yellow crystals of 1 (PF6)2 and orange crystals of 2 (CF3SO3)2 used in the X‐ray analysis were obtained from aqueous solutions of 1 (PF6)2 and 2 (CF3SO3)2. The Rh–Oaqua length of 2.194(4) Å obtained for 1 (PF6)2 is significantly different from that of 2.157(3) Å obtained for the previously reported disorder model [Cp*RhIII(bpy)(0.7H2O/0.3CH3OH)](CF3SO3)2·0.7H2O in which the coordinated water is replaced by a coordinated methanol. The five‐membered ring involving the Rh atom and the 6,6′‐Me2bpy chelating unit in 2 (CF3SO3)2 is not flat, whereas the five‐membered chelate ring in 1 (PF6)2 is nearly flat. Such a non‐planar structure in 2 (CF3SO3)2 is ascribed to the steric repulsion between the 6,6′‐Me2bpy ligand and the Cp* ligand. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

19.
Crystal structures are reported for four (2,2′‐bipyridyl)(ferrocenyl)boronium derivatives, namely (2,2′‐bipyridyl)(ethenyl)(ferrocenyl)boronium hexafluoridophosphate, [Fe(C5H5)(C17H15BN2)]PF6, (Ib), (2,2′‐bipyridyl)(tert‐butylamino)(ferrocenyl)boronium bromide, [Fe(C5H5)(C19H22BN3)]Br, (IIa), (2,2′‐bipyridyl)(ferrocenyl)(4‐methoxyphenylamino)boronium hexafluoridophosphate acetonitrile hemisolvate, [Fe(C5H5)(C22H20BN3O)]PF6·0.5CH3CN, (IIIb), and 1,1′‐bis[(2,2′‐bipyridyl)(cyanomethyl)boronium]ferrocene bis(hexafluoridophosphate), [Fe(C17H14BN3)2](PF6)2, (IVb). The asymmetric unit of (IIIb) contains two independent cations with very similar conformations. The B atom has a distorted tetrahedral coordination in all four structures. The cyclopentadienyl rings of (Ib), (IIa) and (IIIb) are approximately eclipsed, while a bisecting conformation is found for (IVb). The N—H groups of (IIa) and (IIIb) are shielded by the ferrocenyl and tert‐butyl or phenyl groups and are therefore not involved in hydrogen bonding. The B—N(amine) bond lengths are shortened by delocalization of π‐electrons. In the cations with an amine substituent at boron, the B—N(bipyridyl) bonds are 0.035 (3) Å longer than in the cations with a methylene C atom bonded to boron. A similar lengthening of the B—N(bipyridyl) bonds is found in a survey of related cations with an oxy group attached to the B atom.  相似文献   

20.
The excited state dynamics of Tris(2,2′-bipyridine)ruthenium(II) hexafluorophosphate, [Ru(bpy)3(PF6)2], was investigated on the surface of bare and sensitized TiO2 and ZrO2 films. The organic dyes LEG4 and MKA253 were selected as sensitizers. A Stern–Volmer plot of LEG4-sensitized TiO2 substrates with a spin-coated [Ru(bpy)3(PF6)2] layer on top shows considerable quenching of the emission of the latter. Interestingly, time-resolved emission spectroscopy reveals the presence of a fast-decay time component (25±5 ns), which is absent when the anatase TiO2 semiconductor is replaced by ZrO2. It should be specified that the positive redox potential of the ruthenium complex prevents electron transfer from the [Ru(bpy)3(PF6)2] ground state into the oxidized sensitizer. Therefore, we speculate that the fast-decay time component observed stems from excited-state electron transfer from [Ru(bpy)3(PF6)2] to the oxidized sensitizer. Solid-state dye sensitized solar cells (ssDSSCs) employing MKA253 and LEG4 dyes, with [Ru(bpy)3(PF6)2] as a hole-transporting material (HTM), exhibit 1.2 % and 1.1 % power conversion efficiency, respectively. This result illustrates the possibility of the hypothesized excited-state electron transfer.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号